You are on page 1of 33

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/221925642

Textile Organic Dyes � Characteristics, Polluting Effects and


Separation/Elimination Procedures from Industrial Effluents
� A Critical Overview

Chapter · February 2012


DOI: 10.5772/32373 · Source: InTech

CITATIONS READS
204 4,935

2 authors:

Carmen Zaharia Daniela Suteu


Gheorghe Asachi Technical University of Iasi 114 PUBLICATIONS   1,229 CITATIONS   
246 PUBLICATIONS   1,370 CITATIONS   
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Pollution Control Chemistry View project

Environmental advanced technologies in textile industry and integrated surveillance and prevention systems for residual wastewaters,
Contract Parteneriat nr.31-053/2007-2010, P 53/2741 (acronim: ECOTEXENV) View project

All content following this page was uploaded by Carmen Zaharia on 27 October 2014.

The user has requested enhancement of the downloaded file.


3

Textile Organic Dyes –


Characteristics, Polluting Effects and
Separation/Elimination Procedures from
Industrial Effluents – A Critical Overview
Zaharia Carmen and Suteu Daniela
‘Gheorghe Asachi’ Technical University of Iasi,
Faculty of Chemical Engineering and Environmental Protection,
Romania

1. Introduction
The residual dyes from different sources (e.g., textile industries, paper and pulp industries,
dye and dye intermediates industries, pharmaceutical industries, tannery, and Kraft
bleaching industries, etc.) are considered a wide variety of organic pollutants introduced
into the natural water resources or wastewater treatment systems.
One of the main sources with severe pollution problems worldwide is the textile industry
and its dye-containing wastewaters (i.e. 10,000 different textile dyes with an estimated
annual production of 7.105 metric tonnes are commercially available worldwide; 30% of
these dyes are used in excess of 1,000 tonnes per annum, and 90% of the textile products are
used at the level of 100 tonnes per annum or less) (Baban et al., 2010; Robinson et al., 2001;
Soloman et al., 2009). 10-25% of textile dyes are lost during the dyeing process, and 2-20%
are directly discharged as aqueous effluents in different environmental components.
In particular, the discharge of dye-containing effluents into the water environment is
undesirable, not only because of their colour, but also because many of dyes released and
their breakdown products are toxic, carcinogenic or mutagenic to life forms mainly because
of carcinogens, such as benzidine, naphthalene and other aromatic compounds (Suteu et al.,
2009; Zaharia et al., 2009). Without adequate treatment these dyes can remain in the
environment for a long period of time. For instance, the half-life of hydrolysed Reactive Blue
19 is about 46 years at pH 7 and 25°C (Hao et al., 2000).
In addition to the aforementioned problems, the textile industry consumes large amounts of
potable and industrial water (Tables 1, 2 and Fig. 1) as processing water (90-94%) and a
relatively low percentage as cooling water (6-10%) (in comparison with the chemical
industry where only 20% is used as process water and the rest for cooling). The recycling of
treated wastewater has been recommended due to the high levels of contamination in
dyeing and finishing processes (i.e. dyes and their breakdown products, pigments, dye
intermediates, auxiliary chemicals and heavy metals, etc.) (Tables 3, 4 and 5) (adapted from
Bertea A. and Bertea A.P., 2008; Bisschops and Spanjers, 2003; Correia et al., 1994; Orhon et
al., 2001).
Organic Pollutants Ten Years After
56 the Stockholm Convention – Environmental and Analytical Update

Type of finishing process Water consumption, 10-3 m3 /kg textile product


Minimum Medium Maximum
Raw wool washing 4.2 11.7 77.6
Wool finishing 110.9 283.6 657.2
Fabric finishing
• Short process 12.5 78.4 275.2
• Complex processing 10.8 86.7 276.9
Cloth finishing
• Simplified processing 8.3 135.9 392.8
• Complex process 20 83.4 377.8
• Panty processing 5.8 69.2 289.4
Carpet finishing 8.3 46.7 162.6
Fibre finishing 3.3 100.1 557.1
Non-fabrics finishing 2.5 40 82.6
Yarn finishing 33.4 212.7 930.7
Table 1. Specific water consumption in textile finishing processes (adapted from Bertea A. &
Bertea A.P., 2008)

Fig. 1. Specific water consumption in different operations of textile finishing (EPA, 1997)

Operation/Process Water consumption (% from total Organic load (% from total


consumption of the textile plant) organic load of the textile plant)
Minimum Medium Maximum Minimum Medium Maximum
General facilities 6 14 33 0.1 2 8
Preparation 16 36 54 45 61 77
Dyeing 4 29 53 4 23 47
Printing 42 55 38 42 59 75
Wetting 0.3 0.4 0.6 0 0.1 0.1
Fabrics washing 3 28 52 1 13 25
Finishing 0.3 2 4 0.1 3 7
Table 2. Water consumption and organic load in different textile finishing steps (EWA, 2005)
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 57

The most common textile processing technology consists of desizing, scouring, bleaching,
mercerizing and dyeing processes (EPA, 1997):
• Sizing is the first preparation step, in which sizing agents such as starch, polyvinyl
alcohol (PVA) and carboxymethyl cellulose are added to provide strength to the fibres
and minimize breakage.
• Desizing is used to remove sizing materials prior to weaving.
• Scouring removes impurities from the fibres by using alkali solution (commonly sodium
hydroxide) to breakdown natural oils, fats, waxes and surfactants, as well as to emulsify
and suspend impurities in the scouring bath.
• Bleaching is the step used to remove unwanted colour from the fibers by using
chemicals such as sodium hypochlorite and hydrogen peroxide.
• Mercerising is a continuous chemical process used to increase dye-ability, lustre and
fibre appearance. In this step a concentrated alkaline solution is applied and an acid
solution washes the fibres before the dyeing step.
• Dyeing is the process of adding colour to the fibres, which normally requires large
volumes of water not only in the dye bath, but also during the rinsing step. Depending
on the dyeing process, many chemicals like metals, salts, surfactants, organic processing
aids, sulphide and formaldehyde, may be added to improve dye adsorption onto the
fibres.
In general, the textile industry uses a large quantity of chemicals such as:
• Detergents and caustic, which are used to remove dirt, grit, oils, and waxes. Bleaching is
used to improve whiteness and brightness.
• Sizing agents, which are added to improve weaving.
• Oils, which are added to improve spinning and knitting.
• Latex and glues, which are used as binders.
• Dyes, fixing agents, and many in-organics, which are used to provide the brilliant array of
colours the market demands.
• A wide variety of special chemicals, which are used such as softeners, stain release agents,
and wetting agents.
Many of these chemicals become part of the final product whereas the rest are removed
from the fabric, and are purged in the textile effluent.

Type of finished textile product Dyes, g/kg Auxiliaries, g/kg Basic chemical
textile textile product compounds, g/kg
product textile product
Polyester fibres 18 129 126
Fabrics from synthetic fibres 52 113 280
Fabrics from cotton 18 100 570
Dyed fabrics from cellulose fibres 11 183 200
Printed fabrics from cellulose fibres 88 180 807
Table 3. Principal pollutants of textile wastewaters (EWA, 2005)
The annual estimated load with pollutants of a textile wastewater is of: 200,000-250,000 t
salts; 50,000-100,000 t impurities of natural fibres (including biocids) and associated
materials (lignin, sericine, etc.); 80,000-100,000 t blinding agents (especially starch and its
derivatives, but also polyacrylates, polyvinyl alcohol, carboxymethyl cellulose); 25,000-
Organic Pollutants Ten Years After
58 the Stockholm Convention – Environmental and Analytical Update

30,000 t preparation agents (in principal, mineral oils); 20,000-25,000 t tensides (dispersing
agents, emulsifiers, detergents and wetting agents); 15,000-20,000 tonnes carboxylic acids
(especially acetic acid); 10,000-15,000 t binders; 5,000-10,000 t urea; 5,000-10,000 t ligands,
and < 5,000 t auxiliaries (EWA, 2005). The environmental authorities have begun to target
the textile industry to clean up the wastewater that is discharged. The principal quality
indicators that regulators are looking for polluting effect or toxicity are the high salt content,
high Total Solids (TS), high Total Dissolved Solids (TDS), high Total Suspended Solids
(TSS), Biological Oxygen Demand (BOD) and Chemical Oxygen Demand (COD), heavy
metals, colour of the textile effluent (ADMI color value - American Dye Manufacturer
Institute color value), and other potential hazardous or dangerous organic compounds
included into each textile processing technological steps (Tables 4 and 5).

Process Textile effluent


Singering, Desizing High BOD, high TS, neutral pH
Scouring High BOD, high TS, high alkalinity, high temperature
Bleaching, Mercerizing High BOD, high TS, alkaline wastewater
Heat-setting Low BOD, low solids, alkaline wastewater
Dyeing, Printing & Wasted dyes, high BOD, COD, solids, neutral to alkaline
Finishing wastewater
Table 4. Wet processes producing textile wastewater (adapted from Naveed S. et al., 2006)

Process COD, BOD, TS, TDS, pH Colour Water usage,


g O2/L g O2/L g/L g/L (ADMI) L/kg product
Desizing 4.6-5.9 1.7-5.2 16.0-32.0 - - - 3-9
Scouring 8.0 0.1-2.9 7.6-17.4 - 10-13 694 26-43
Bleaching 6.7-13.5 0.1-1.7 2.3-14.4 4.8-19.5 8.5-9.6 153 3-124
Mercerising 1.6 0.05-0.1 0.6-1.9 4.3-4.6 5.5-9.5 - 232-308
Dyeing 1.1-4.6 0.01-1.8 0.5-14.1 0.05 5-10 1450-4750 8-300
Table 5. Principal characteristics of a cotton wet processing wastewater (Cooper, 1995)
The wastewater composition is depending on the different organic-based compounds,
inorganic chemicals and dyes used in the industrial dry and wet-processing steps. Textile
effluents from the dyeing and rinsing steps represent the most coloured fraction of textile
wastewaters, and are characterized by extreme fluctuations in many quality indicators such
as COD, BOD, pH, colour, salinity and temperature.
The colour of textile wastewater is mainly due to the presence of textile dyes, pigments and
other coloured compounds. A single dyeing operation can use a number of dyes from
different chemical classes resulting in a complex wastewater (Correia et al., 1994). Moreover,
the textile dyes have complex structures, synthetic origin and recalcitrant nature, which
makes them obligatory to remove from industrial effluents before being disposed into
hydrological systems (Anjaneyulu et al, 2005).
The dye removal from textile effluent is always connected with the decolourization
treatment applied for textile wastewater in terms of respectation the local environmental
quality requirements and standards (Table 6) (i.e. removal values of COD, BOD, TS, TSS,
TDS, colour, total nitrogen, and total phosphorus from textile wastewater higher than 70-
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 59

85% or concentration values of the specific quality indicators under the imposed or standard
limits) (Zaharia, 2008).
The decolorization treatments applied for different textile effluents include current and also
advanced non-biological (i.e. specific mechano-physical, chemical, electrochemical process-
es, etc.) and also biological processes (Suteu et al., 2009; Zaharia, 2006; Zaharia et al, 2011).

Quality indicator M.A.C.*, mg/L


Discharge directly
Discharge in urban Water bodies quality,
in water bodies
WW sewerage class I -natural non-
network polluted state
pH 6.5-8.5 6.5-8.5 6.5-8.5
BOD5 25 300 3
COD 125 500 10
TSS 35 350 -
TDS 2000 250 < 500
Total N 10 15 1.5
Total P 1 5 0.1
Sulphates, SO42- 600 600 80
Chlorides, Cl- 500 500 <100
Sulphides (S2-) + H2S 0.5 1.0 <0.5
Synthetic detergents 0.5 25 <0.5
Others (Oil & grease) 20 30 <0.1
* M.A.C. – Maximum Admissible Concentration

Table 6. Romanian national wastewater and water quality standard adapted to European
and international standards (adapted from Zaharia, 2008)
Studies on the behavior of textile organic dyes in water and wastewater treatment
processes refer predominantly to laboratory tests or investigations of semi-technical
plants, sometimes under conditions related to waterworks practice. In addition, textile
operators, water supply companies, local environmental authorities have collected a lot of
data on the behavior of textile dyes during textile wastewater treatment, but have seldom
published their results.
However, information on the behavior of textile organic dyes is needed, because the limited
number of reports available that are based on realistic operating conditions or which
reproduce practical conditions that are already several years old.

2. Textile organic dyes – Classification and characteristics


The dyes are natural and synthetic compounds that make the world more beautiful through
coloured products. The textile dyes represent a category of organic compounds, generally
considered as pollutants, presented into wastewaters resulting mainly from processes of
chemical textile finishing (Suteu et al., 2011a; Zaharia et al., 2009).
The textile coloration industry is characterised by a very large number of dispersed
dyehouses of small and medium size that use a very wide range of textile dyes.
Organic Pollutants Ten Years After
60 the Stockholm Convention – Environmental and Analytical Update

2.1 Textile organic dye classification


The nature and origin are firstly considered as criteria for the general classification in
natural and synthetic textile dyes.
The natural textile dyes were mainly used in textile processing until 1856, beginning in 2600
BC when was mentioned the use of dyestuff in China, based on dyes extracted from
vegetable and animal resources. It is also known that Phoenicians were used Tyrian purple
produced from certain species of crushed sea snails in the 15th century BC, and indigo dye
produced from the well-known indigo plant since 3000 BC. The dyes from madder plants
were used for wrapping and dyeing of Egyptian mummies clothes and also of Incas fine
textures in South America.
The synthetic dyes were firstly discovered in 1856, beginning with ‚mauve‘ dye (aniline),
a brilliant fuchsia colour synthesed by W.H. Perkin (UK), and some azo dyes synthesed by
diazotisation reaction discovered in 1958 by P. Gries (Germany) (Welham, 2000). These
dyes are aromatic compounds produced by chemical synthesis, and having into their
structure aromatic rings that contain delocated electrons and also different functional
groups. Their color is due to the chromogene-chromophore structure (acceptor of
electrons), and the dyeing capacity is due to auxochrome groups (donor of electrons). The
chromogene is constituted from an aromatic structure normally based on rings of
benzene, naphthaline or antracene, from which are binding chromofores that contain
double conjugated links with delocated electrons. The chromofore configurations are
represented by the azo group (-N=N-), ethylene group (=C=C=), methine group (-CH=),
carbonyl group (=C=O), carbon-nitrogen (=C=NH; -CH=N-), carbon-sulphur (=C=S; ≡C-
S-S-C≡), nitro (-NO2; -NO-OH), nitrozo (-N=O; =N-OH) or chinoid groups. The
auxochrome groups are ionizable groups, that confer to the dyes the binding capacity
onto the textile material. The usual auxochrome groups are: -NH2 (amino), -COOH
(carboxyl), -SO3H (sulphonate) and -OH (hydroxyl) (Suteu et al, 2011; Welham, 2000). Five
examples of textile dyes are presented in Fig. 2.
The textile dyes are mainly classified in two different ways: (1) based on its application
characteristics (i.e. CI Generic Name such as acid, basic, direct, disperse, mordant, reactive,
sulphur dye, pigment, vat, azo insoluble), and (2) based on its chemical structure
respectively (i.e. CI Constitution Number such as nitro, azo, carotenoid, diphenylmethane,
xanthene, acridine, quinoline, indamine, sulphur, amino- and hydroxy ketone,
anthraquinone, indigoid, phthalocyanine, inorganic pigment, etc.) (Tables 7 and 8).
Excepting the colorant precursors such as azoic component, oxidation bases and sulphur
dyes, almost two-third of all organic dyes are azo dyes (R1-N=N-R2) used in a number of
different industrial processes such as textile dyeing and printing, colour photography,
finishing processing of leather, pharmaceutical, cosmetics, etc. The starting material or
intermediates for dye production are aniline, chloroanilines, naphthylamines,
methylanilines, benzidines, phenylenediamines, and others.
Considering only the general structure, the textile dyes are also classified in anionic,
nonionic and cationic dyes. The major anionic dyes are the direct, acid and reactive dyes
(Robinson et al., 2001), and the most problematic ones are the brightly coloured, water
soluble reactive and acid dyes (they can not be removed through conventional treatment
systems).
The major nonionic dyes are disperse dyes that does not ionised in the aqueous
environment, and the major cationic dyes are the azo basic, anthraquinone disperse and
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 61

reactive dyes, etc. The most problematic dyes are those which are made from known
carcinogens such as benzidine and other aromatic compounds (i.e. anthroquinone-based
dyes are resistant to degradation due to their fused aromatic ring structure). Some disperse
dyes have good ability to bioaccumulation, and the azo and nitro compounds are reduced in
sediments, other dye-accumulating substrates to toxic amines (e.g. R1-N=N-R2 + 4H+ + 4e- →
R1-NH2 + R2-NH2).
The organic dyes used in the textile dyeing process must have a high chemical and
photolytic stability, and the conventional textile effluent treatment in aerobic conditions
does not degrade these textile dyes, and are presented in high quantities into the natural
water resources in absence of some tertiary treatments.

Reactive Orange 16
OH
O C.I. 18097
N
a
O
S
O
C
H
-
C
H
-
S
O
3

N=N N C Anionic monoazo reactive dye;


CH 3
NaO 3S MW = 617.54 g/mol;
λmax = 495 nm
Cl Cl Brilliant Red HE-3B
SO3Na N N N N NaSO3 (Reactive Red 120)
OH HN N N N
C.I. 25810
N H H N NH OH
N N
N Anionic, bifunctional
azo reactive dye;
NaSO3 SO3Na NaSO3 SO3Na MW = 1463 g/mol;
λmax = 530 nm
Crystal Violet
(CH3)2N C N+(CH3)2 Cl
- (Basic Violet 3)
C.I. 42555
Cationic triphenylmethane dye;
MW = 407.99 g/mol;
N(CH3)2
λmax = 590 nm
+ +
N(C2H5)2
- Rhodamine B
(C2H5)2N O Cl
(Basic Violet 10)
C
C.I. 45170
COOH
Cationic, Xantenic dye;
MW =479.2 g/mol;
λmax = 550 nm

Methylene Blue
N (Basic Blue 9)
C.I. 52015
+
S
Cationic, phenothiazine dye;
N(CH3)2
(CH3)2N Cl- MW =319.85 g/mol;
λmax= 660 nm

Fig. 2. Chemical structure and principal characteristics of different textile dyes


(Suteu et al., 2011a)
Organic Pollutants Ten Years After
62 the Stockholm Convention – Environmental and Analytical Update

Chemical class C.I. Constitution Chemical class C.I. Constitution


numbers numbers
Nitroso 10000-10299 Indamine 49400-49699
Nitro 10300-10099 Indophenol 49700-49999
Monoazo 11000-19999 Azine 50000-50999
Disazo 20000-29999 Oxazine 51000-51999
Triazo 30000-34999 Thiazine 52000-52999
Polyazo 35000-36999 Sulphur 53000-54999
Azoic 37000-39999 Lactone 55000-56999
Stilbene 40000-40799 Aminoketone 56000-56999
Carotenoid 40800-40999 Hydroxyketone 57000-57999
Diphethylmethane 41000-41999 Anthraquinone 58000-72999
Triarylmethane 42000-44999 Indigoid 73000-73999
Xanthene 45000-45999 Phthalocyanine 74000-74999
Acridine 46000-46999 Natural 75000-75999
Quinoline 47000-47999 Oxidation base 76000-76999
Methine 48000-48999 Inorganic pigment 77000-77999
Thiazole 49000-49399
Table 7. Colour index classification of dye chemical constituents (Cooper, 1995)

Chemical class Distribution between application ranges, %


Acid Basic Direct Disperse Mordant Pigment Reactive Solvent Vat
Unmetallised azo 20 5 30 12 12 6 10 5 -
Metal complex 65 - 10 - - - 12 13 -
Thiazole - 5 95 - - - - - -
Stilbene - 2 98 - - - - - -
Anthraquinone 15 2 - 25 3 4 6 9 36
Indigoid 2 - - - - 17 - - 81
Quinophthalene 30 20 - 40 - - 10 - -
Aminoketone 11 - - 40 8 - 3 8 20
Phtalocyanine 14 4 8 - 4 9 43 15 3
Formazan 70 - - - - - 30 - -
Methine - 71 - 23 - 1 - 5 -
Nitro, nitroso 31 2 - 48 2 5 - 12 -
Triarylmethane 35 22 1 1 24 5 - 12 -
Xanthene 33 16 - - 9 2 2 38 -
Acridine - 92 - 4 - - - 4 -
Azine 39 39 - - - 3 - 19 -
Oxazine - 22 17 2 40 9 10 - -
Thiazine - 55 - - 10 - - 10 25
Table 8. Distribution of each chemical class between major application ranges (adapted from
Cooper, 1995)
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 63

The major textile dyes can be included in the two high classes: azo or anthraquinone (65-
75% from total textile dyes). The azo dyes are characterised by reactive groups that form
covalent bonds with HO-, HN-, or HS- groups in fibres (cotton, wool, silk, nylon). Azo dyes
are mostly used for yellow, orange and red colours. Anthraquinone dyes constitute the
second most important class of textile dyes, after azo dyes, and have a wide range of colours
in almost the whole visible spectrum, but they are most commonly used for violet, blue and
green colours (Fontenot et al., 2003). Considering the nature of textile fibres that are dyeing,
the textile dyes can be classified as into Table 9.

Class Subclass PES CA PAN PA Silk Wool Cotton


Disperse +++ +++ ++ ++ - - -
Cationic - ∼ +++ ++ - - -
Standard - - ∼ +++ +++ +++ -
Acid 1:1 - - - P + +++ -
1:2 - - - ++ + +++ -
Reactive - - - ∼ ++ ++ +++
Direct - - - ++ ++ P +++
Stuff ∼ - ∼ ∼ ∼ − +++
Indigoid - - − ∼ ∼ ∼ P
Sulphur - + − − − − +++
Insoluble azo +++ - ∼ ∼ ∼ − +++
Legend: +++ very frequent; ++ frequent; + sometimes; ∼ possible; P especially printing
Table 9. Dye classes and dyeing textile substrates (adapted from Bertea A. & Bertea A.P., 2008)
The textile dyeing process is due to physico-chemical interactions developed at contacting of
textile material with dye solution or dispersion, which contains a large variety of chemicals
(salts, acids) and dyeing auxiliaries (tensides, dispersing agents, etc.).

2.2 Textile dye characterisation


The identification of individual unknown dyes in a coloured effluent or watercourse is
difficult to be done and implies advanced analytical methods (i.e. individual and/or
coupled spectrophotometry, G/L chromatography and mass spectrometry procedures), and
also the colour determination and appreciation in different operating situations.
The characterisation and identification data of the textile dyes as main chemicals in dyeing
process must consist of:
• dye identity data (i.e. name, C.I. or CAS number, molecular and structural formula;
composition, degree of purity, spectral data; methods of detection and determination)
(e.g., some examples illustrated in Fig. 2),
• dye production information (i.e. production process, proposed uses, form,
concentration in commercially available preparations, estimated production,
recommended methods and precautions concerning handling, storage, transport, fire
and other dangers, emergency measures, etc.) (e.g. some indications in Table 3),
• dye physico-chemical properties (i.e. boiling point (b.p.), relative density, water
solubility, partition coefficient, vapour pressure, self-ignition, oxidising properties,
granulometry, particle size distribution, etc.) (e.g., for the first synthetic discovered dye:
Aniline – 184 (b.p.), kH= 2.05E-01, Csatw= 3.6E+04, pKa= 4.6, log Kow= 0.90; for 4,4’-
Organic Pollutants Ten Years After
64 the Stockholm Convention – Environmental and Analytical Update

Methylenedianiline - 398 (b.p.), kH= 5.67E-06, Csatw= 1000, pKa= n.s., log Kow= 1.59; for 4-
Aminodiphenylamine - 354 (b.p.), kH= 3.76E-01, Csatw= 1450, pKa= 5.2, log Kow= 1.82;
where, kH - the Henry’s constant at 1013 hPa and 25°C (Pa.m3/mol), Csatw - the water
solubility (mg/L) at 25°C, pKa – dissociation constant of the protonated azo dye at 25°C,
Kow – n-octanol/water partition coefficient, n.s. – not specified)
• toxicological studies (i.e. acute toxicity-oral, inhalation, dermal, skin or eyes irritation,
skin sensitisation, repeated dose toxicity-28 days, mutagenicity, toxicity to
reproduction, toxicokinetic behaviour),
• ecotoxicological studies (i.e. acute toxicity to fish, daphnia, growth inhibition on algae,
bacteriological inhibition, degradation: ready biodegradability, abiotic degradation-
hydrolysis as a function of pH, BOD, COD, BOD/COD ratio).
These characteristics and indentification data are given obligatory by the dye producers or
distributors of textile products on the free market of homologated colorants, and also exist
in the library data of some operating programs of advanced analysis apparatus.
The textile azo dyes are characterized by relatively high polarity (log Kow up to 3) and high
recalcitrance. Recalcitrance is difficult to evaluate because of the dependence of degradation
on highly variable boundary conditions (e.g., redox milieu or pH). For example, aniline (the
first synthetic discovered dye) is known to be easily degradable, but under specific anoxic
conditions it has been proven to be easily stable (Börnick and Schmidt, 2006). Furthermore,
the azo dyes are relevant in terms of eco- and human toxicity, industrially produced in high
quantities, and known to occur in hydrosphere.
The azo dyes can accept protons because of the free electron pair of the nitrogen, and the
free electron pair of nitrogen interacts with the delocalized π-orbital system.
Acceptor substituients at the aromatic ring such as –Cl or –NO2 cause an additional decrease
in the basic character of aminic groups. Donor groups such as –CH3 or –OR (in metha and
para position) lead to an increase in the basicity of aromatic aminic groups. However, donor
substituients in the ortho position can sterically impede the protonation and consequently
decrease the basicity of aminic groups. The azo dyes are characterized by amphoteric
properties when molecules contain additional acidic groups such as hydroxyl, carboxyl or
sulfoxyl substituents.
Depending on pH value, the azo dyes can be anionic (deprotonation at the acidic group),
cationic (protonated at the amino group) or non-ionic. Accordingly, knowledge of the
acidity constants is indispensable for the characterization of the behavior of azo dyes.
Environmental partitioning is influenced by substituents as well as the number of carbon
atoms and aromatic structure of the carbon skeleton. The presence of an amino group causes
a higher boiling point, a higher water solubility, a lower Henry’s law constant, and a higher
mobility in comparison with hydrocarbons (the amino group can also reduce the mobility
by specific interactions with solids via covalent bonding to carbonyl moieties or cation
exchange) (Börnick and Schmidt, 2006). The volatility of azo dyes in aqueous solution is in
most cases very low. Some colour characteristics from different studies, tests and literature
data are presented in Table 10, especially for reactive dyes.
The chromophore distribution in reactive dyes indicated that the great majority of
unmetallised azo dyes are yellow, orange and red. Contrarily, the blue, green, black and
brown contain a much more proportion of metal-complex azo, anthraquinone,
triphenyldioxazine or copper phthalocyanine chromophores (Table 10).
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 65

Chemical class Yellow Orange Red Violet Blue Green Brown Black Proportion
of all
reactive
dyes
Unmetallised azo 97 90 90 63 20 16 57 42 66
Metal complex azo 2 10 9 32 17 5 43 55 15
Anthraquinone - - - 5 34 37 - 3 10
Phthalocyanine - - - - 27 42 - - 8
Miscellanous 1 - 1 - 2 - - - 1
Table 10. Distribution of chemical classes in reactive dye range (adapted from Cooper, 1995)
In general, colour in wastewater is classified in terms of true/real colour (i.e. colour of
turbidity-free water sample), or apparent colour (i.e. colour of non-treated water sample).
The most common methods to measure the colour of dye solution or dispersion, and/or
wastewater are visual comparison and spectrophotometry, although there is still a lack of an
universal method to classify coloured wastewater discharges. By visual comparison, colour is
quantified by comparing the colour of sample with either known concentrations of coloured
standards (normally a platinum-cobalt solution), or properly calibrated colour disks, and is
less applicable for highly coloured industrial wastewaters. In the spectrophotometric method,
colour-measuring protocols differ between the methodologies, of which the most commonly
used are Tristimulus Filter Method, American Dye Manufacturer Institute (ADMI)
Tristimulus Filter Method, and Spectra record (Table 11).

Spectrophotometric Description
method
Tristimulus Three tristimulus light filters combined with a specific light source
(i.e. tungsten lamp) and a photoelectric cell inside a filter
photometer. The output transmittance is converted to trichromatic
coefficient and colour characteristic value.
ADMI Tristimulus The ADMI colour value provides a true watercolour measure,
which can be differentiated in 3 (WL) ADMI (i.e. the transmittance
is recorded at 590, 540 and 438 nm) or 31 (WL) ADMI (i.e. the
transmittance is determined each 10 nm in the range of 400-700 nm).
Spectra record The complete spectrum is recorded, and the entire spectrum, or a
part of it, is used for comparison. A modified method has been
suggested in which areas beneath an extinction curve represent
the colour intensity, being expressed as space units.
Table 11. Spectrophotometric methods for colour determination in dye solution or
dispersion, water and wastewater (adapted from Dos Santos et al., 2004).

2.3 Dye fixation on textile fibres


In general, textile fibres can catched dyes in their structures as a result of van der Waals
forces, hydrogen bonds and hydrophobic interactions (physical adsorption). The uptake of
the dye in fibres depends on the dye nature and its chemical constituents. But the strongest
dye-fibre attachment is a result of a covalent bond with an additional electrostatic
interaction where the dye ion and fibre have opposite charges (chemisorption).
Organic Pollutants Ten Years After
66 the Stockholm Convention – Environmental and Analytical Update

In alkaline conditions (i.e. pH 9-12), at high temperatures (30-70ºC), and salt concentration
from 40-100 g/L, reactive dyes form a reactive vinyl sulfone (─SO3─CH═CH2) group, which
creates a bond with the fibres. However, the vinyl sulfone group undergoes hydrolysis (i.e. a
spontaneous reaction that occurs in the presence of water), and because the products do not
have any affinity with the fibres, they do not form a covalent bond (Dos Santos et al., 2004).
Therefore, a high amount of dye constituents are discharged in the wastewater.
The fixation efficiency varies with the class of azo dye used, which is around 98% for basic
dyes and 50% for reactive dyes (Table 12) (Bertea A. & Bertea A.P., 2008; O'Neill et al., 1999).
Large amounts of salts such as sodium nitrate, sodium sulphate and sodium chloride are
used in the dyebath, as well as sodium hydroxide that is widely applied to increase the pH
to the alkaline range. It is estimated that during the mercerising process the weight of these
salts can make up 20% of the fibre weight (EPA, 1997).

Dye class Fibre type Fixation degree, % Loss in effluent, %


Acid Polyamide 80-95 5-20
Basic Acrilic 95-100 0-5
Direct Cellulose 70-95 5-30
Disperse Polyester 90-100 0-10
Metal complex Wool 90-98 2-10
Reactive Cellulose 50-90 10-50
Sulphur Cellulose 60-90 10-40
Dye-stuff Cellulose 80-95 5-20
Table 12. Fixation degree of different dye classes on textile support (EWA, 2005).
The problem of high coloured effluent or dye-containing effluent has become identified
particularly with the dyeing of cellulose fibres (cotton – 50% of the total consumed fibres in
the textile industry worldwide), and in particular with the use of reactive dyes (10-50% loss
in effluent), direct dyes (5-30% loss in effluent), vat dyes (5-20% loss in effluent), and
sulphur dyes (10-40% loss in effluent).
The research for dynamic response and improved dyeing productivity has served to focus
the attention of the textile coloration industry on right-first-time production techniques that
minimise wastes, make important contribution to reduce colour loads in the effluent by
optimisation of processes, minimising of dye wastage, and control automatically the dyeing
and printing operation.
After the textile dyeing and finishing processes, a predicted environmental concentration of a
dye in the receiving water can be estimated based on the following factors: (i) daily dye
usage; (ii) dye fixation degree on the substrate (i.e. textile fibres or fabrics); (iii) dye
removal degree into the effluent treatment process, and (iv) dilution factor in the receiving
water.
Some scenario analyses were mentioned the values of dye concentration in some receiving
rivers of 5-10 mg/L (average value, 50 days each year) or 1300-1555 mg/L (the worst case, 2
days each year) for batchwise dyeing of cotton with reactive dyes, and of 1.2-3 mg/L
(average value, 25 days each year) or 300-364 mg/L (the worst case, 2 days each year) for
batchwise dyeing of wool yarn with acid dyes (adapted from Cooper Ed., 1995).
Limits on dye-containing organic loads will become more restrictive in the future, which
makes cleaning exhausts an environmental necessity.
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 67

3. Textile organic dyes – Environmental problems and polluting effects


The environmental issues associated with residual dye content or residual colour in treated
textile effluents are always a concern for each textile operator that directly discharges, both
sewage treatment works and commercial textile operations, in terms of respecting the colour
and residual dye requirements placed on treated effluent discharge (Zaharia et al., 2011).
Dye concentrations in watercourses higher of 1 mg/L caused by the direct discharges of
textile effluents, treated or not, can give rise to public compliant. High concentrations of
textile dyes in water bodies stop the reoxygenation capacity of the receiving water and cut-
off sunlight, thereby upsetting biological activity in aquatic life and also the photosynthesis
process of aquatic plants or algae (Zaharia et al., 2009).
The colour in watercourses is accepted as an aesthetic problem rather than an eco-toxic
hazard. Therefore, the public seems to accept blue, green or brown colour of rivers but the
‘non-natural’ colour as red and purple usually cause most concern.
The polluting effects of dyes against aquatic environment can be also the result of toxic
effects due to their long time presence in environment (i.e. half-life time of several years),
accumulation in sediments but especially in fishes or other aquatic life forms, decomposition
of pollutants in carcinogenic or mutagenic compounds but also low aerobic
biodegradability. Due to their synthetic nature and structure mainly aromatic, the most of
dyes are non-biodegradable, having carcinogenic action or causing allergies, dermatitis, skin
irritation or different tissular changes. Moreover, various azo dyes, mainly aromatic
compounds, show both acute and chronic toxicity. High potential health risk is caused by
adsorption of azo dyes and their breakdown products (toxic amines) through the
gastrointestinal tract, skin, lungs, and also formation of hemoglobin adducts and
disturbance of blood formation. LD50 values reported for aromatic azo dyes range between
100 and 2000 mg/kg body weight (Börnick & Schmidt, 2006).
Several azo dyes cause damage of DNA that can lead to the genesis of malignant tumors.
Electron-donating substituents in ortho and para position can increase the carcinogenic
potential. The toxicity diminished essentially with the protonation of aminic groups. Some
of the best known azo dyes (e.g. Direct Black 38 azo dye, precursor for benzidine;
azodisalicylate, precursor for 4-phenylenediamine) and their breakdown derivatives
inducing cancer in humans and animals are benzidine and its derivatives, and also a large
number of anilines (e.g. 2-nitroaniline, 4-chloroaniline, 4,4’-dimethylendianiline, 4-
phenylenediamine, etc.), nitrosamines, dimethylamines, etc.
The main pollution characteristics and category (pollution risk) of the principal products
used in processing of textile materials are summarized in Table 13 (EWA, 2005).
In different toxicological studies are indicated that 98% of dyes has a lethal concentration
value (LC50) for fishes higher than 1 mg/L, and 59% have an LC50 value higher than 100
mg/L (i.e. 31% of 100-500 mg/L and 28% higher than 500 mg/L).
Other ecotoxicological studies indicated that over 18% of 200 dyes tested in England showed
significant inhibition of the respiration rate of the biomass (i.e. wastewater bacteria) from
sewage, and these were all basic dyes (adapted from Cooper, 1995).
The bioaccumulation potential of dyes in fish was also an important measure estimating the
bioconcentration factor (dye concentration in fish/dye concentration in water). No
bioaccumulation is expected for dyes with solubility in water higher than 2000 mg/L.
Dyes are not biodegradable in aerobic wastewater treatment processes and some of them
may be intactly adsorbed by the sludge at wastewater biological treatment (i.e.
bioelimination by adsorptive removal of dyes).
Organic Pollutants Ten Years After
68 the Stockholm Convention – Environmental and Analytical Update

Products used in textile industry Pollution characteristics Pollution


category
Alkali, mineral acids, salts, oxidants Inorganic pollutants, 1
relatively inofensive
Sizing agents based on starch, natural Easy biodegradables;
oils, fats, waxes, biodegradable with a moderate - high 2
surfactants, organic acids, reducing BOD5
agents
Colorants and optic whitening agents, Difficult to be 3
fibres and impurities of polymeric biodegraded
nature, synthetic polymeric resins,
silicones
Polyvinyl alcohols, mineral oils, tensides Difficult to be 4
resistant to biodegradation, anionic or biodegraded; moderate
non-ionic emolients BOD5
Formaldehyde or N-methylolic reagents, Can not be removed by
coloured compounds or accelerators, conventional biological 5
retarders and cationic emolients, treatment, low BOD5
complexants, salts of heavy metals
Table 13. Characterization of products used in textile industry vs. their polluting effect
Some investigations of adsorption degree onto sludge of some azo dyes indicated typically
high levels of adsorption for basic or direct dyes, and high to medium range for disperse
dyes, all the others having very low adsorption, which appears to depend on the
sulphonation degree or ease of hydrolysis. These tested azo dyes are presented in Table 14.

Group 1 Group 2 Group 3


Dyes unaffected by biological treatment Dyes eliminated by Dyes with high
adsorption on sludge biodegradability
CI Acid Yellow 17 CI Acid Red CI Acid Red 151 CI Acid Orange 7
CI Acid Yellow 23 CI Acid Red 14 CI Acid Blue 113 CI Acid Orange 8
CI Acid Yellow 49 CI Acid Red 18 CI Direct Yellow 28 CI Acid Red 88
CI Acid Yellow 151 CI Acid Red 337 CI Direct Violet 9
CI Acid Orange 10 CI Acid Black 1
CI Direct Yellow 4
Table 14. Fate of water-soluble azo dyes in the activated sludge treatment (Cooper, 1995)
The high degree of sulphonation of azo dyes in Group 1 enhanced their water solubility and
limited their ability to be adsorbed on the biomass. The dyes in Group 2 were also highly
sulphonated but permitted a relatively good adsorption performance on sludge.
Other information in bioelimination of different reactive dyes mentioned that monoazo dyes
are particularly poorly adsorbed, and disazos, anthraquinones, triphendioxazines and
phthalocyanines are generally much better adsorbed than monoazos.
It is important to underline that toxic compounds (e.g. toxic aromatic amines, benzidine and
its derivatives) can be formed in the environment via transformation of textile dye-
precursors (e.g., reduction or hydrolysis of textile azo dyes). The textile dye-precursors are
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 69

introduced in water environment due to industrial production of dyes and industrial


production of textile fibres, fabrics and clothes via wastewater, sludge, or solid deposits.
The quality problem of dye content and/or colour in the dyehouse effluent discharged in
watercourses can be solved by using of a range of advanced decolourisation technologies
investigated by the major dye suppliers, textile operators and customers who are under
pressure to reduce colour and residual dye levels in their effluents.

4. Textile organic dyes – Separation and elimination procedures from water


environment (especially industrial wastewater)
The textile organic dyes must be separated and eliminated (if necessary) from water but
especially from industrial wastewaters by effective and viable treatments at sewage
treatment works or on site following two different treatment concepts as: (1) separation of
organic pollutants from water environment, or (2) the partial or complete mineralization or
decomposition of organic pollutants. Separation processes are based on fluid mechanics
(sedimentation, centrifugation, filtration and flotation) or on synthetic membranes (micro-,
ultra- and nanofiltration, reverse osmosis). Additionally, physico-chemical processes (i.e.
adsorption, chemical precipitation, coagulation-flocculation, and ionic exchange) can be
used to separate dissolved, emulsified and solid-separating compounds from water
environment (Anjaneyulu et al., 2005; Babu et al., 2007; Robinson et al., 2001; Suteu et al.,
2009a; Suteu et al., 2011a; Zaharia, 2006; Zaharia et al., 2009; Zaharia et al., 2011).
The partial and complete mineralization or decomposition of pollutants can be achieved by
biological and chemical processes (biological processes in connection with the activated
sludge processes and membrane bioreactors, advanced oxidation with ozone, H2O2, UV)
(Dos Santos et al., 2004 ; Oztekin et al., 2010 ; Wiesmann et al., 2007 ; Zaharia et al., 2009).
A textile operator will decide on options available to plan forward strategy that will ensure
compliance with the environmental regulators’ requirements on a progressive basis focused
on some options and applied solutions of different separation processes (sedimentation,
filtration, membrane separation), and some physico-chemical treatment steps (i.e.
adsorption; coagulation-flocculation with inorganic coagulants and organic polymers;
chemical oxidation; ozonation; electrochemical process, etc.) integrated into a specific order
in the technological process of wastewater treatment for decolourization or large-scale
colour and dye removal processes of textile effluents.
To introduce a logical order in the description of treatment methods for textile dye and
colour removal, the relationship between pollutant and respective typical treatment
technology is taken as reference. The first treatment step for textile wastewater and also
rainwater is the separation of suspended solids and immiscible liquids from the main textile
effluents by gravity separation (e.g., grit separation, sedimentation including
coagulation/flocculation), filtration, membrane filtration (MF, UF), air flotation, and/or
other oil/water separation operations.
The following treatment steps are applied to soluble pollutants, when these are transferred
into solids (e.g., chemical precipitation, coagulation/flocculation, etc.) or gaseous and
soluble compounds with low or high dangerous/toxic effect (e.g., chemical oxidation,
ozonation, wet air oxidation, adsorption, ion exchange, stripping, nanofiltration/reverse
osmosis). Solid-free wastewater can either be segregated into a biodegradable and a non-
biodegradable part, or the contaminants responsible for the non-biodegradable wastewater
part that can be decomposed based on physical and/or chemical processes. After an
Organic Pollutants Ten Years After
70 the Stockholm Convention – Environmental and Analytical Update

adequate treatment, the treated wastewater (WW) can either be discharged into a receiving
water body, into a subsequent central biological wastewater treatment plant (BWWTP) or a
municipal wastewater treatment plant (MWWTP).
Some selected treatment processes for dyes and colour removal of industrial wastewater
applied over the time into different textile units are summarized in Table 15. Some of these
methods will be further detail and some of authors’ results summarized.

Treatment Treatment Advantages Limitations


methodology stage
Physico-chemical treatments
Precipitation, Pre/main Short detention time and low Agglomerates
coagulation- treatment capital costs. Relatively good separation and
flocculation removal efficiencies. treatment. Selected
operating condition.
Electrokinetic Pre/main Economically feasible High sludge production
coagulation treatment
Fenton process Pre/main Effective for both soluble andSludge generation;
treatment insoluble coloured problem with sludge
contaminants. No alternation in
disposal. Prohibitively
volume. expensive.
Ozonation Main Effective for azo dye removal.Not suitable for
treatment Applied in gaseous state: no dispersed dyes.
alteration of volume Releases aromatic dyes.
Short half-life of ozone
(20 min)
Oxidation with Post treatment Low temperature requirement. Cost intensive process.
NaOCl Initiates and accelerates azo- Release of aromatic
bond cleavage amines
Adsorption with solid adsorbents such as:
Activated carbon Pre/post Economically attractive. Good Very expensive; cost
treatment removal efficiency of wide intensive regeneration
variety of dyes. process
Peat Pre treatment Effective adsorbent due to Surface area is lower
cellular structure. No activation than activated carbon
required.
Coal ashes Pre treatment Economically attractive. Good Larger contact times and
removal efficiency. huge quantities are
required. Specific
surface area for
adsorption are lower
than activated carbon
Wood chips/ Pre treatment Effective adsorbent due to Long retention times
Wood sawdust cellular structure. Economically and huge quantities are
attractive. Good adsorption required.
capacity for acid dyes
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 71

Silica gels Pre treatment Effective for basic dyes Side reactions prevent
commercial application
Irradiation Post treatment Effective oxidation at lab scale Requires a lot of
dissolved oxygen (O2)
Photochemical Post treatment No sludge production Formation of by-
process products
Electrochemical Pre treatment No additional chemicals Cost intensive process;
oxidation required and the end products mainly high cost of
are non-dangerous/hazardous. electricity
Ion exchange Main Regeneration with low loss of Specific application; not
treatment adsorbents effective for all dyes
Biological treatments
Aerobic process Post treatment Partial or complete Expensive treatment
decolourization for all classes
of dyes
Anaerobic Main Resistant to wide variety of Longer acclimatization
process treatment complex coloured compounds. phase
Bio gas produced is used for
stream generation.
Single cell Post treatment Good removal efficiency for Culture maintenance is
(Fungal, Algal & low volumes and cost intensive. Cannot
Bacterial) concentrations. Very effective cope up with large
for specific colour removal. volumes of WW.
Emerging treatments
Other advanced Main Complete mineralization Cost intensive process
oxidation process treatment ensured. Growing number of
commercial applications.
Effective pre-treatment
methodology in integrated
systems and enhances
biodegradability.
Membrane Main Removes all dye types; High running cost.
filtration treatment recovery and reuse of chemicals Concentrated sludge
and water. production. Dissolved
solids are not separated
in this process
Photocatalysis Post treatment Process carried out at ambient Effective for small
conditions. Inputs are no toxic amount of coloured
and inexpensive. Complete compounds. Expensive
mineralization with shorter process.
detention times.
Sonication Pre treatment Simplicity in use. Very effective Relatively new method
in integrated systems. and awaiting full scale
application.
Enzymatic Post treatment Effective for specifically Enzyme isolation and
treatment selected compounds. purification is tedious.
Organic Pollutants Ten Years After
72 the Stockholm Convention – Environmental and Analytical Update

Unaffected by shock loadings Efficiency curtailed due


and shorter contact times to the presence of
required. interferences.
Redox mediators Pre/ Easily available and enhances Concentration of redox
supportive the process by increasing mediator may give
treatment electron transfer efficiency antagonistic effect. Also
depends on biological
activity of the system.
Engineered Pre/post Cost effective technology and High initial installation
wetland systems treatment can be operated with huge cost. Requires expertise
volumes of wastewater and managing during
monsoon becomes
difficult
Table 15. Various current and emerging dye separation and elimination treatments applied
for textile effluents with their principal advantages and limitations (adapted from
Anjaneyulu et al., 2005; Babu et al., 2007; Robinson et al., 2001)

4.1 Physical treatments


4.1.1 Adsorption
One of the most effective and proven treatment with potential application in textile
wastewater treatment is adsorption. This process consists in the transfer of soluble organic
dyes (solutes) from wastewater to the surface of solid, highly porous, particles (the
adsorbent). The adsorbent has a finite capacity for each compound to be removed, and when
is ‘spent’ must be replaced by fresh material (the ‘spent’ adsorbent must be either
regenerated or incinerated).
Adsorption is an economically feasible process for dyes removal and/or decolourization of
textile effluents being the result of two mechanisms: adsorption and ion exchange. The
principal influencing factors in dye adsorption are: dye/adsorbent interaction, adsorbent
surface area, particle size, temperature, pH, and contact time. Adsorbents which contain
amino nitrogen tend to have a significantly larger adsorption capacity in acid dyes.
The most used adsorbent is activated carbon, and also other commercial inorganic
adsorbents. Some ‘low cost’ adsorbents of industrial or agricultural wastes (i.e. peat, coal
ashes, refused derived coal fuel, clay, bentonite and modified bentonite, red soil, bauxite,
ebark, rice husk, tree barks, neem leaf powder, wood chips, ground nut shell powder, rice
hulls, bagasse pith, wood sawdust, grounded sunflower seed shells, other ligno-cellulosic
wastes, etc.) are also used for removal of dye and organic coloured matter from textile
effluents (i.e. a removal of 40-90% basic dyes and 40% direct dyes, with maximum
adsorption capacities for basic dyes of 338 mg/g) (Anjaneyulu et al., 2005; Bhattacharyya &
Sarma, 2003; Gupta et al., 1992; Nigam et al., 2000; Ozcan et al., 2004; Robinson et al., 2001;
Suteu & Zaharia, 2008; Suteu et al., 2009b; Suteu et al., 2011a,b; Zaharia et al., 2011). The use
of these materials is advantageous mainly due to their widespread availability and
cheapness. Sometimes the regeneration is not necessary and the ‘spent’ material is
conventionally burnt although there is potential for solid state fermentation (SSF) for protein
enrichment. The use of ‘low cost’ adsorbents for textile dye removal is profitable but
requires huge quantity of adsorbents, being lower efficient than activated carbon. Some
authors’ results for dye adsorption are summarized in Table 16.
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 73

Adsorptive Operating conditions / Optimal doses Adsorption efficiency for


material (Adsorption + Sedimentation/Filtration) some tested textile dyes, %
Peat pH= 2 (BRed); 5.7 (MB, RhB); Cdye= (20-300) (67.50 – 85.60) BRed
mg/L BRed; (19-134) mg/L MB; (65.70– 89.30) MB
(28-155) mg/L RhB; Cadsorbant= 12 g/L (98.30– 99.00 RhB
Wood pH= 2 (BRed); 5,7 (MB; CV; RhB); 1 (RO); (63.90– 83.00) BRed
sawdust Cdye= (20-150) mg/L BRed; (6-40) mg/L MB; (69.40 – 91.80) MB
(8-50) mg/L CV; (9-58) mg/L RhB; (24-160) (66.00 – 80.00) CV
mg/L (RO); Cadsorbant= 20 g/L (BRed); (52.20 – 72.00) RhB
4 g/L (MB, CV, RhB); 8 g/L (RO) (38.30 – 50.80) RO
Sunflower pH= 1 (RO); 6 (MB); Cdye= (25–160) mg/L RO; (80.60 – 82.60) RO
seed shell (25–280) mg/L MB; Cadsorbant= 8 g/L (RO), (84.40 – 93.00) MB
4 g/L (MB)
Corn cobs pH= 1 (RO) or 6 (MB); Cdye= (24-160) mg/L (83.00 – 86.20) RO
RO; (25-280) mg/L MB; Cadsorbant= 8 g/L (RO), (87.20 – 98.10) MB
4 g/L (MB)
Lignine pH= 1.5 (BRed); 1 (RO); 6 (MB); Cdye= (50-300) (49.50-59.50) BRed
mg/L BRed; (25.60-280) mg/L MB; (30-150) (43.50-60.70) RO
mg/L RO; Cadsorbant= 14 g/L (BRed); (55.30-64.30) MB
4 g/L (MB); 12 g/L (RO)
Cellolignine pH=6; Cdye= (25.6-281.6) mg/L MB; (96.60 – 98.40) MB
Cadsorbant= 4 g/L
Abbreviations: BRed - Brilliant Red HE-3B (Reactive Red 120)/CI 25810; RO - Reactive
Orange 16/CI 17757; MB - Methylene Blue (Basic Blue 9)/CI 52015; RhB - Rhodamine B
(Basic Violet 10)/CI 45170; CV - Crystal Violet (Basic Violet 3)/CI 42555
Table 16. Dye adsorption performance onto some natural adsorptive materials (adapted
from Suteu et al., 2009b; Suteu et al., 2011a,b).
Adsorption with activated carbon. Activated carbon have been engineered for optimal
adsorption of the contaminants found in dyehouse effluents: large, negatively charged or
polar molecules of dyes. Powdered or granular activated carbon (specific surface area of
500-1500 m2/g; pore volume of 0.3-1 cm3/g; bulk density of 300-550 g/L) has a reasonably
good colour removal capacity when is introduced in a separate filtration step. The activated
carbon is used as granulate (GAC) in columns or as powder (PAC) in batchwise treatment
into a specific treatment tank or basin. High removal rates are obtained for cationic mordant
and acid dyes (Anjaneyulu et al., 2005), and a slightly lesser extent (moderate) for dispersed,
direct, vat, pigment and reactive dyes (Cooper, 1995; Nigam et al., 2000) with consumable
doses of 0.5-1.0 kg adsorbent/m3 wastewater (i.e. dye removal of 60-90%).
Most recent studies mentioned that an effective irreversible adsorption of dye molecules
onto the adsorbent particles takes place via a combination of physical adsorption of dye
onto adsorbent surfaces within the microporous structure of the particles, enhanced by an
ion-exchange process wherein the interlayer anions of the adsorbent are displaced by the
dye molecules, and also inter-particle diffusion processes. Removal of pollutants can take
place at any pH between 2 and 11, and at any temperature between 0 and 100°C (effluent
temperature as received, generally 30-40°C).
Organic Pollutants Ten Years After
74 the Stockholm Convention – Environmental and Analytical Update

The adsorption on activated carbon without pretreatment is imposible because the


suspended solids rapidly clog the filter, and may be feasible in combination with
flocculation-sedimentation treatment or a biological treatment (Masui et al., 2005; Ramesh
Babu et al., 2007). The main important disadvantage of this process is attributed to the high
cost of activated carbon. Performance is dependent on the type of activated carbon used and
wastewater characteristics, and can be well suited for one particular wastewater system and
ineffective in another. The activated carbon has to be reactivated otherwise disposal of the
concentrates has to be considered (reactivation results in 10-15% loss of the adsorbent).

4.1.2 Irradiation
The irradiation treatment is a simple and efficient procedure for eliminating a wide variety
of organic contaminants, and as well disinfecting harmful microorganism using gamma rays
or electron beams (e.g., source for irradiation can be a monochromatic UV lamps working
under 253.7 nm). A high quantity of dissolved oxygen is required for an organic dye to be
effectively broken down by irradiation. The dissolved oxygen is consumed very rapidly and
so a constant and adequate supply is required. Irradiation treatment of a secondary effluent
from sewage treatment plant reduced COD, TOC and colour up to 64%, 34% and 88%
respectively, at a dose of 15 K Gy gamma-rays (Borrely et al., 1998). The efficiency of
irradiation treatment increases when is used catalyst as titanium dioxide (Krapfenbauer et
al., 1999). A lot of data are reported with the practical results obtained at the simple
exposure of different dye solutions or dispersions and dye-containing textile wastewaters to
sunlight for a period of a half, one or two months (direct photolysis with natural sunlight
into open basins). All these reports indicated high removals of colour (>84%), dye
destruction by photooxidation following first order kinetics at treatment of some vat dye
effluents. But the direct photolysis of textile organic dye in the natural aquatic environment
has proven difficult due to strong dependence of the decay rates on dye reactivity and
photosensitivity. Most of all commercial dyes are usually designed to be light resistant.
Therefore, the recent researches have been directed towards investigation of organic dye
photodegradation by sensitizers or catalysts in aqueous/dispersion systems by UV
irradiation. Moreover, there are reported high removal of indigo-colour when is initiated a
laser fading process for indigo coloured denim textile mainly based on basic interaction of
laser beam with indigo-coloured textile (Dascalu et al., 2000).

4.1.3 Membrane processes


The increasing of water cost and necessity of reduction of water consumption implies
treatment process which is integrated with in-plant water circuits rather than subsequent
treatment (Baban et al., 2010; Machenbach, 1998). From this point of view, membrane
filtration offers potential application in combination with other textile effluent treatments.
Membrane processes for wastewater treatment are pressure-driven processes, capable to
clarify, concentrate, and most important, separate dye discontinuously from effluent (Xu &
Lebran, 1991). These are new technologies, which can restrict organic contaminants and
microorganisms presented in wastewater (i.e. color removal, BOD reduction, salt reduction,
Polyvinyl Acetate (PVA) recovery, and latex recovery). The common membrane filtration
types are: Micro-Filtration (MF), Ultra-Filtration (UF), Nano-Filtration (NF), and Reverse
Osmosis (RO). The choice of the membrane process must be guided by the required quality
of the final effluent.
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 75

Micro-filtration is mainly used for treatment of dye baths containing pigment dyes as well
as for subsequent rinsing baths (Ramesh Babu et al., 2007). Chemicals that can not be
filtrated by microfiltration will remain in the dye bath. Microfiltration can be used as a
pretreatment for nanofiltration or reverse osmosis (Ghayeni et al., 1998), and also to separate
suspended solids, colloids from effluents or macromolecules with pores of 0.1 to 1 micron.
MF performance is typically of >90% for turbidity or silt density index. Microfiltration
membranes are made of specific polymers such as Poly (Ether Sulfone), Poly (Vinylidiene
Fluoride), Poly (Sulfone), Poly (Vinylidene Difluoride), Polycarbonate, Polypropylene, Poly
Tetrafluoroethylene (PTFE), etc. Ceramic, glass, carbon, zirconia coated carbon, alumina and
sintered metal membranes have been employed where extraordinary chemical resistance or
where high temperature operation is necessary. MF and UF operate at 20 to 100 psi
transmembrane pressures (Ptm) (low pressure membrane procees) and velocities of 20 to
100 cm/s (Naveed et al., 2006).
Ultra-filtration is used to separate macromolecules and particles, but the elimination of
polluting substances such as dyes is never complete (only 31-76% dye removal). The quality
of treated wastewater does not permit its reuse for sensitive processes, such as textile dyeing
(Ramesh Babu et al., 2007) but permit recycling of 40% treated wastewater in stages in which
salinity is not a problem, such as rinsing, washing, etc. Ultrafiltration can only be used as a
pretreatment for reverse osmosis (Ciardelli & Ranieri, 2001) or in combination with a
biological reactor (Mignani et al., 1999) or to remove metal hydroxides (reducing the heavy
metal content to 1 ppm or less) (Naveed et al., 2006). UF membranes are made of polymeric
materials (i.e. polysulfone, polypropylene, nylon-6, polytetrafluoroethylene (PTFE),
polyvinyl chlorides (PVC), acrylic copolymer etc.
Nano-Filtration was used for the treatment of coloured effluents from the textile industry,
mainly in a combination of adsorption (for decreasing of concentration polarization during
the filtration process) and nanofiltration (NF modules are extremely sensitive to fouling by
colloidal material and macromolecules). NF membranes are generally made of cellulose
acetate and aromatic polyamides, and retain low-molecular weight organic compounds,
divalent ions, large monovalent ions, hydrolized reactive dyes, and dyeing auxiliaries.
Inorganic materials, such as ceramics, carbon based membranes, zirconia, are also used in
manufacturing NF and RO membranes. Typical NF flux rates are 5 to 30 GFD (Gross Flow
per Day) (Naveed et al., 2006). A performance of above 70% colour removal for a NF plant
was reported working at 8 bar/18°C, with four polyethersulphonate membranes with
molecular weight cut offs of 40, 10, 5 and 3 kda for three different effluents coming from
dyeing cycle of textile industry (Alves & Pinho, 2000). Values of colour removal higher than
90% were reported for single NF process, and also combination MF and NF, in the case of
different effluents from textile fabrics processing. Harmful effects of high concentrations of
dye and salts in the dye house effluents were frequently reported (i.e. concentration of dye >
1.5 g/L, and of mineral salts >20 g/L) (Tang & Chen, 2002). An important problem is the
acculumation of dissolved solids, which makes discharge of treated effluents in
watercourses almost impossible. NF treatment can be an alternative fairly satisfactory for
textile effluent decolourization.
Reverse Osmosis is used to remove in a single step most types of ionic compounds,
hydrolized reactive dyes, chemical auxiliaries, and produce a high quality of permeate
(Ramesh Babu et al., 2007). Like NF, RO is very sensitive to fouling and the influent must be
carefully pretreated. RO membranes are generally made of cellulose acetate and aromatic
polyamides but also of inorganic materials. The Ptm in RO is typically 500 to 1000 psi, with
Organic Pollutants Ten Years After
76 the Stockholm Convention – Environmental and Analytical Update

cross flows of 20 to 100 cm/s. The range of typical RO fluxes is 5 to 15 GFD (Naveed et al.,
2006). In combination with physio-chemical treatment, the membrane processes has
advantages over the other conventional treatments, such as the ability to recover materials
with valuable recyclable water, reducting fresh water consumption and wastewater
treatment costs, small disposal volumes which minimizes waste disposal costs, reduction of
regulatory pressure and fine improved heat recovery systems. Membrane processes have
many cost-effective applications in textile industry.

4.2 Chemical treatment


4.2.1 Oxidative processes
Chemical oxidation represents the conversion or transformation of pollutants by chemical
oxidation agents other than oxygen/air or bacteria to similar but less harmful or hazardous
compounds and/or to short-chained and easily biodegradable organic components
(aromatic rings cleavage of dye molecules).
The modern textile dyes are resistant to mild oxidation conditions such those existing in
biological treatment systems. Therefore, efficient dye and colour removal must be
accomplished by more powerful oxidising agents such as chlorines, ozone, Fenton reagents,
UV/peroxide, UV/ozone, or other oxidising procedures or combinations.
Oxidative processes with hydrogen peroxide. The oxidation processes with hydrogen
peroxide (H2O2) (oxidation potential, E°= 1.80 V at pH 0, and E°= 0.87 V at pH 14) can be
explored as wastewater treatment alternatives in two systems: (1) homogenous systems
based on the use of visible or ultraviolet light, soluble catalysts (Fenton reagents) and other
chemical activators (e.g. ozone, peroxidase etc.) and (2) heterogenous systems based on the
use of semiconductors, zeolites, clays with or without ultraviolet light, such as TiO2, stable
modified zeolites with iron and aluminium (i.e. FeY5, FeY11.5 etc.) (difficulty encountered in
the separation of the solid photocatalysts at the end of the process) (Neamtu et al., 2004;
Zaharia et al., 2009).
Fenton reagent is usually hydrogen peroxide (H2O2) that is activated by some iron salts (i.e.
Fe2+ salts) (without UV irradiation) to form hydroxyl radicals (HO.) which are strong
oxidants (oxidation potential, E°= 3.06 V) than H2O2 and ozone. The Fenton oxidation
reactions are detailed in other chapters of this book, and the treatment efficiency depends
mainly of effluent characteristics, and operating parameters (e.g., colour removal of 31.10 or
56.20%, at a pH of 4.00, for Fenton oxidation of textile Remazol Arancio 3R, Remazol Rose
RB dye-containing effluents working with 0.18-0.35 M H2O2 and 1.45 mM Fe2+, after 30 or
120 min) (Zaharia et al., 2011).
Heterogenous catalytic oxidation with 20 mM H2O2 and FeY11.5 (1 g/L) of Procion Marine H-
EXL dye-containing effluents lead to colour removal of 53-83%, COD removal of 68-76% and
TOC removal of 32-37% at pH=3-5, after 10 min of oxidation (Neamtu et al., 2004). Working
with FeY5 (1 g/L) and 20 mM H2O2, at pH=3 and 5, for the same textile effluent the treatment
efficiency, was of 95 and 35% for colour and COD removal after 10 min of oxidation, and 97%
for colour after 60 min of oxidation; COD removal (60 min) of 64.20% (Zaharia, 2006).
When small quantities of wastewater are involved or when there is no biotreatment
available at the textile site, chemical oxidation might be recommendable treatment option
instead of installing a central biological WWT plant. Advantages of this oxidative treatment
include reduction of effluent COD, colour and toxicity, and also the possibility to be used to
remove both soluble and insoluble dyes (i.e. disperse dyes). Complete decolourization was
obtained after the complete Fenton reagent stage (generally 24 hours).
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 77

Ozonation process. Ozone is a powerful oxidising agent (oxidation potential, E°= 2.07 V)
capable of cleavage the aromatic rings of some textile dyes and descomposition of other
organic pollutants from industrial effluents. The ozone decomposes the organic dyes with
conjugated double bonds forming smaller molecules with increased carcinogenic or toxic
properties, and so ozonation may be used alongside a physical method to prevent this (i.e.
irradiation, membrane separation, adsorption, etc). Ozone can react directly or indirectly
with dye molecules. In the direct pathway, the ozone molecule is itself the electron acceptor,
and hydroxide ions (i.e. pH > 7-8) catalyze the auto decomposition of ozone to hydroxyl
radicals (·OH) in aqueous effluents (very strong and non-selective oxidants) which react
with organic and inorganic chemicals. At low pH ozone efficiently reacts with unsaturated
chromophoric bonds of a dye molecule via direct reactions (Adams & Gorg, 2002).
The main advantage is that ozone can be applied in its gaseous state and therefore does not
increase the volume of wastewater and sludge. A disadvantage of ozonation is its short half-
life, tipically being 20 min, the destabilisation by the presence of salts, pH, and temperature,
and the additional costs for the installation of ozonation plant. The improvement of
ozonation preformance is obtained in combination with irradiation (Surpateanu & Zaharia,
2004a; Zaharia et al., 2009) or with a membrane filtration technique (Lopez et al., 1999).
Treatment of dye-containing wastewater with ozone followed by chemical coagulation
using Ca(OH)2 indicated 62% colour removal after ozonation (Sarasa et al., 1998).
Oxidation process with sodium hypochlorite. This treatment implies the attack at the
amino group of the dye molecule by Cl+, initiating and accelerating azo-bond cleavage. The
increasing of chlorine concentration favors the dye removal and decolourization process,
and also the decreasing of pH. The dye containing amino or substituted amino groups on
the naphthalene ring (i.e. dyes derived from amino-naphtol- and naphtylamino-sulphonic
acids) are most susceptible for chlorine decolourization (Omura, 1994). This treatment is
unsuitable for disperse dyes, and is becoming less frequent due to the negative effects at
releasing into watercourses of aromatic amines or otherwise toxic molecules. Moreover,
althought about 40% of the pigments used worldwise contain chlorine this corresponds to
only less than 0.02% of the total chlorine production (Slokar & Le Marechal, 1997).
Photochemical oxidation process. The UV treatment in the presence of H2O2 can
descomposed dye molecules to low weight organic molecules, or even to CO2, H2O, other
inorganic oxides, hydrides, etc. There can be also produced additional by-products such as
halides, metals, inorganic acids, organic aldehydes and organic acids depending on initial
materials and the extent of decolourisation treatment (Yang et al., 1998). The dye
decomposition is initiated by the generated hydroxyl radicals (H2O2 + hυ → 2HO·) and
hydroperoxide radicals (H2O2 + HO· → HO2· + H2O).
The treatment may be set-up in a batch or continuous column unit, and is influenced by the
intensity of the UV radiation, pH, dye structure and the dye bath composition (Slokar & Le
Marechal, 1997). The performance of photooxidation treatment in the presence of hydrogen
peroxide are high (i.e. >60-90% for colour removal, working with 400-500 mg/L H2O2 at pH
3-7, for Red M5B, H-acid and Blue MR dye-containing effluents) (Anjaneyulu et al., 2005) or
81-94% dye removal after 60 min, working with 88 mM H2O2 at pH of 4-6, for Acid Red G
dye-containing effluent (Surpateanu & Zaharia, 2004b; Zaharia et al., 2009).
Electrochemical oxidation process. As an advanced process, the electrochemical treatment
of dye-containing effluents is a potentially powerful method of pollution control, offering
high removal efficiencies (Anjaneyulu et al., 2005) especially for acid dyes as well as
Organic Pollutants Ten Years After
78 the Stockholm Convention – Environmental and Analytical Update

disperse and metal complex dyes. The main advantages of this treatment are considered the
requirement of simple equipment and operation, low temperature in comparison with other
non-electrochemical treatments, no requirement of any additional chemicals, easy control
but crucial for pH, the electrochemical reactors (with electrolytic cells) are compact, and
prevent the production of unwated by-products. The principal oxidising agent in
electrochemical process is hypochlorite ion or hypochlorous acid produced from naturally
occuring chloride ions. Hydroxyl radical and other reactive species also participate in
electrochemical oxidation of organics (Kim et al., 2002) that can be achieved directly or
indirectly at the anode. The breakdown compounds are generally not hazardous being
discharged into watercourses without important environmental and health risks.
The electrochemical oxidation is considered an efficient and economic treatment of recycling
textile wastewater for the dyeing stage. The environmental advantage mainly achieved is
the minimization of all emissions: emission of gases, solid waste, and liquid effluent. Other
important advantage is its capacity of adaptation to different volumes and pollution loads.
The main disadvantage is the generation of metallic hydroxide sludge (from the metallic
electrodes in the cell), that limits its use (Ramesh Babu et al., 2007).
Some studies reported colour removals of about 100% for dyeing wastewater within only 6
min of electrolysis (Vlyssides et al., 2000) (e.g., complete decolourization of textile effluent
containing blue-26 anthraquinone dye by electrochemical oxidation with lead dioxide
coated anode - Titanium Substract Insoluble Anode (TSIA), at neutral pH, in the presence of
sodium chloride, current density of 4.5 A/dm2, electrolysis time of 220 min, or maximum
95.2% colour and 72.5% COD removal of textile azo dye-containing effluent in a flow reactor
working at rate of 5 mL/min and current density of 29.9 mA/cm2) (Anjaneyulu et al., 2005).

4.2.2 Coagulation-flocculation and precipitation


It is clearly known that the coloured colloid particles from textile effluents cannot be
separated by simple gravitational means, and some chemicals (e.g., ferrous sulphate, ferric
sulphate, ferric chloride, lime, polyaluminium chloride, polyaluminium sulphate, cationic
organic polymers, etc.) are added to cause the solids to settle. These chemicals cause
destabilisation of colloidal and small suspended particles (e.g. dyes, clay, heavy metals,
organic solids, oil in wastewater) and emulsions entrapping solids (coagulation) and/or the
agglomeration of these particles to flocs large enough to settle (flocculation) or highly
improve further filtration (Zaharia et al., 2006; Zaharia et al., 2007). In the case of
flocculation, anionic and non-ionic polymers are also used.
The mechanism by which synthetic organic polymer removes dissolved residual dyes from
effluents is best described in terms of the electrostatic attraction between the oppositely
charged soluble dye and polymer molecules. Many of the most problematic dye types, such
as reactive dyes, carry a residual negative charge in their hydrolysed dissolved form, and so
positively charged groups on the polymers provide the neccesary counter for the interaction
and subsequent precipitation to occur. The immediate result of this coprecipitation is the
almost instantaneous production of very small coloured particles, having little strength and
breaking down at any significant disturbances. The agglomeration of the coloured
precipitates by using appropriate high polyelectrolyte flocculants produces stable flocs
(Zaharia et al., 2007, 2011). The main disadvantages of this treatment are the process control
that is a little difficult, the potential affection of precipitation rate and floc size by impurities
such as non-ionic detergents remaining in the effluent, and the sludge production which has
to be settled, dewatered and pressed into a cake for subsequent landfilling tipping.
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 79

There are reported very effective chemical coagulation-flocculation (C-F) and


precipitation of textile wastewater which reduced the load on the biological treatment,
working with polyaluminium chloride along with an organic polymer (Lin & Chen, 1997)
or ferrous/ferric chloride and a commercial organic coagulant aid at pH of 6.7-8.3 (colour
removal > 80%) (Venkat Mohan et al., 1999) or alum at pH=8.2 (54-81% colour removal)
with addition of bentonite (3 g/L) for Remazol Violet dye-containing effluent (Sanghi et
al., 2001). Other efficient textile treatments mentioned by different textile operators consist
in coagulation-flocculation followed by membrane technology (especially for recycling
textile effluents).
Some of authors’ results in different (C-F) treatments are summarized in Table 17.

Process Coagulant/ Wastewater characteristics or dye type / Results


Flocculant
Coagulation- Ferrous sulphate (5 Wastewater characteristics: pH=6.5-7; TSS=250-1000
flocculation/ mg/L) + Ponilit GT-2 mg/L; colour=650 UH; CODCr=152.7-272 mg O2/L
sedimentation/ anionic polyelectrolyte Process efficiency: turbidity removal=70.31-91.34%;
filtration (15 mg/L) + bentonite colour removal=70.20-90.50% and CODCr
(3 g/L) removal=34.90-45.20%
Precipitation/ NaOH + Na2CO3 + Wastewater characteristics: pH=7.0-9.5; total metallic
flocculation/ Ca(OH)2, Ponilit GT-2 ions= 48.30 mg/L; extractable substances in
flotation with anionic polyelectrolyte organic solvents= 980 mg/L
dissolved air Process efficiency: Metallic ions removal=78.67-92.33%
Coagulation- Ferric sulfate (2-5 Wastewater characteristics: pH=6.98; T=20°C,
flocculation/ mg/L) +Prodefloc Turbidity=556 FTU; CODCr=152.60 mg O2/L;
sedimentation/ CRC 301 (0.25-1.5 colour=1320 UH
filtration mg/L) cationic Process efficiency: maximal turbidity
polyelectrolyte removal=95.87% and colour removal= 93.90-98.10%
Table 17. Some applications of coagulation- flocculation in wastewater treatment (adapted
from Zaharia, 2006; Zaharia et al., 2006, 2007).

4.2.3 Electrocoagulation
An advanced electrochemical treatment for dye and colour removal is electrocoagulation
(EC) that has as main goal to form flocs of metal hydroxides within the effluent to be
cleaned by electro-dissolution of soluble anodes. EC involves important processes as
electrolytic reactions at electrodes, formation of coagulants in aqueous effluent and
adsorption of soluble or colloidal pollutants on coagulants, and removal by sedimentation
and flotation. This treatment is efficient even at high pH for colour and COD removals being
strongly influenced by the current density and duration of reaction. The EC treatment was
applied with high efficiency for textile Orange II and Acid red 14 dye-containing effluents
(i.e. > 98% colour removal) (Daneshvar et al., 2003) or Yellow 86 dye-containing wastewater
(i.e. turbidity, COD, extractible substances, and dye removal of 87.20%, 49.89%, 94.67%, and
74.20%, after 30 min of operation, current intensity of 1 A, with monopolar electrodes)
(Zaharia et al, 2005) when iron is used as sacrificial anode. In general, decolorization
performance in EC treatment is between 90-95%, and COD removal between 30-36% under
optimal conditions (Ramesh Babu et al., 2007).
Organic Pollutants Ten Years After
80 the Stockholm Convention – Environmental and Analytical Update

4.2.4 Ionic exchange


The ion exchange process has not be widely used for treatment of dye-containing effluents,
mainly because of the general opinion that ion exchangers cannot accomodate a wide range
of dyes (Slokar & Le Marechal, 1997).
The ionic exchange occurs mainly based on the interaction of ionic species from wastewater
with an adsorptive solid material, being distinguished from the conventional adsorption by
nature and morphology of adsorptive material or the inorganic structure containing
functional groups capable of ionic exchange (Macoveanu et al., 2002). The mechanisms of
ionic exchange process are well known, and two principal aspects must be mentioned: (1)
ionic exchange can be modeling as well as adsorption onto activated coal; (2) the ion
exchangers can be regenerated without modifying the equilibrium condition (e.g., by
passing of a salt solution containing original active groups under ion exchanger layer). In
the case of wastewater treatment, the effluent is passed over the ion exchanger resin until
the available exchange sites are saturated (both cationic and anionic dyes are removed).
The ionic exchange is a reversible process, and the regenerated ion exchanger can be reused.
The essential characteristic of ionic exchange that makes distinction of adsorption is the fact
that the replace of ions takes place in stoechiometric proportion (Macoveanu et al., 2002).
The effluent treatment by ionic exchange process contributes to the diminishing of energetic
consumption and recovery of valuable components under diverse forms, simultaneously
with the wastewater treatment. In practice, the ion exchangers are used in combination with
other wastewater treatments. The main advantages of ion exchange are removal of soluble
dyes, no loss of adsorbent at regeneration, and reclamation of solvent after use. The
important disadvantages of this process is the cost, organic solvents are expensive, and ion
exchange treatment is not efficient for disperse dyes (Robinson et al., 2001). Our results in
batchwise treatment of Brilliant Red HE-3B dye-containing effluents (0.05 – 0.3 mg/mL)
using anionic Purolite A-400 and Purolite A-500 indicated dye removal of 48-89% working
in the optimal conditions, or dye removal of 56-78% for Crystal Violet (Basic Violet 3) using
Purolite C-100 or dye removal between 78-89% for Reactive Blue M-EB using ion exchange
celluloses (Suteu et al., 2002).

4.3 Biological treatments


Biological treatments are considered reproduction, artificially or otherwise, of self-
purification phenomena existing in natural environment. There are different biological
treatments, performed in aerobic or anaerobic or combined anaerobic/aerobic conditions.
The processing, quality, adaptability of microorganisms, and the reactor type are decisive
parameters for removal efficiency (Börnick & Schmidt, 2006).
Biological treatment process for decolorization of industrial effluents is ambiguous, different
and divergent (Anjaneyulu et al., 2005). Previous subchapters indicate that dyes themselves
are not biologically degradable since microorganisms do not use the coloured constituents
as a source of food. The most currently used biodegradation involve aerobic micro-
organisms, which utilize molecular oxygen as reducing equivalent acceptor during the
respiration process. But biodegradation in anaerobic environment conditions (anoxic and
hypoxic environments) also occurs, and survival of microorganisms is possible by using
sulphates, nitrates and carbon dioxide as electron acceptors (Birch et al., 1989).
Research data indicates that certain dyes are susceptible to anoxic/anaerobic decolouri-
zation, and also that an anaerobic step followed by an aerobic step may represent a
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 81

significant advancement in biological decolourization treatment in future (Ong et al., 2005).


The treatment plant that receives dye-containing effluents has high potential to form toxic
biodegradation products such as toxic amines, benzidine and its derivates, etc. To avoid that
risk, anaerobic/aerobic sequential reactor systems seem to be an efficient procedure (i.e.
efficient colour removal takes place during the anaerobic treatment, and high reduction of
aromatic amines and other organic compounds occurs during the subsequent aerobic
treatment). Under aerobic conditions, most of the azo dye metabolites are quickly degraded
by oxidation of the substituents or of the side branches. However, some of them are still
rather recalcitrant. Successful removal of poorly degradable amines was often achieved by
adaptation of microorganisms (i.e. acclimatization of biological sludge to nitroaniline-
containing wastewaters; after gradual adaptation, the microorganisms are able to eliminate
3- and 4-nitroaniline simultaneously) (Börnick & Schmidt, 2006). Difficult to be biodegraded
are also the aromatic amines containing sulfo substituents in metha position (e.g., 3-
aminobenzenesulfonate) that can be treated with good efficiencies with flow-through
bioreactors within 28 days. Contrarily, some experimental results found anaerobic
mineralization to be efficient for aromatic amines.
The main advantage of biological treatment in comparison with certain physico-chemical
treatments is that over 70% of organic matter expressed by CODCr may be converted to
biosolids (Anjaneyulu et al., 2005).

4.3.1 Aerobic biological treatment


Biological treatment with ‘activated sludge’ was the most used in large scale textile
effluent treatment, and the trickling filter or biological aerated filter (BAF) is an
alternative, permitting a 34-44% dye-colour removal for different high dyeing loads of
industrial effluents. The main microorganisms contributing to biodegradation of organic
compounds are bacteria (e.g. Bacillus subtilis, Aeromonas Hydrophilia, Bacillus cetreus,
Klebsiella pneunomoniae, Acetobacter liquefaciens, Pseudomonas species, Pagmentiphaga kullae,
Sphingomonas, etc.), fungi (e.g., white-rot fungi: Phanerochaete chrysosporium, Hirschioporus
larincinus, Inonotus hispidus, Phlebia tremellosa, Coriolus versicolor, etc.), algae (e.g. Chlorella
and oscillotoria species) etc. Moreover, some bacteria, white-rot fungi, mixed microbial
cultures from a wide variety of habitats are found to be able to degrade dyes using
enzimes, such as lignin peroxidases (LiP), manganese dependent peroxidases (MnP),
H2O2-producing enzime such as glucose-1-oxidase and glucose-2-oxidase, along with
laccase, and a phenoloxidase. Biological aerated filters involve the growth of an organism
on inert media that are held stationary during normal operation and exposed to aeration.
In aerobic conditions, the mono- and dioxygenase enzymes catalyse the incorporation of
dissolved oxygen into the aromatic ring of organic compounds prior to ring fission.
Although azo dyes are aromatic compounds, their nitro and sulfonic groups are quite
recalcitrant to aerobic bacterial degradation (Dos Santos et al., 2004). However, in the
presence of specific oxygen-catalysed enzymes called azo reductases, some aerobic
bacteria are able to reduce azo compounds and produce aromatic amines (Stolz, 2001).
The batch experiments with aerobic activated sludge confirmed the biodegradability of
sulphonated azo dyes. Only aerobic degradation of the azo dyes is possible by azo
reduction (i.e. high colour removal (>90%) of Red RBN azo dye-containing effluents (3000
mg/L) working with Aeromonas hydrophilla in the specific optimal conditions of pH (5.5-
10), temperature (20°-38°C), and time (8 days)), and mineralization does not occur. The
Organic Pollutants Ten Years After
82 the Stockholm Convention – Environmental and Analytical Update

subsequential anaerobic and aerobic bioreactor was able to completely remove the
sulphonated azo dye (i.e. MY10) at a maximum loading rate of 210 mg/L per day (Tan et
al., 2000). The degradation of azo dyes (i.e. Acid Red 151; Basic Blue 41; Basic Red 46, 16;
Basic Yellow 28, 19) in an aerobic biofilm system indicated 80% colour removal
(Anjaneyulu et al., 2005). The improvement of dye biodegradation performance (i.e. >90%
colour removal) are made by adding activated carbon (PAC) or bentonite in aeration tank.

4.3.2 Anaerobic biological treatment


Anaerobic biodegradation of azo and other water-soluble dyes is mainly reported as an
oxidation-reduction reaction with hydrogen, and formation of methane, hydrogen sulphide,
carbon dioxide, other gaseous compounds, and releasing electrons. The electrons react with
the dye reducing the azo bonds, causing the effluent decolourization. Azo dye is considered
an oxidising agent for the reduced flavin nucleotides of the microbial electron chain, and is
reduced and decolourized concurrently with reoxidation of the reduced flavin nucleotides
(Robinson et al., 2001). An additional carbon organic source is necessary, such as glucose
which is a limiting factor in scale set-up technology application. The azo and nitro-
components are reduced in anoxic sediments and in the intestinal environment, with
regeneration of toxic amines (Banat et al., 1996). A major advantage of anaerobic system
along with effluent decolourization is the production of biogas, reusable for heat and power
generation that will reduce energy costs. Since textile industry wastewaters are generally
discharged at high temperatures (40–70°C), thermophilic anaerobic treatment could serve as
an interesting option, especially when closing process water cycles is considered. Anaerobic
decolourization of textile effluents (e.g., colour removal of >99% for a Orange II, Black 2HN
under anaerobic condition, more than 72 h) is not yet well established although successful
pilot-scale and full-scale plants are very well operating (Tan et al., 2000). Among the
different studied reactors, anaerobic filter and UASB thermophilic anaerobic reactor gave
good colour removals, using or not redox mediator (e.g., antraquinone-2,6, disulphonic acid)
as catalyst capable of acceleration the colour removal of azo dye-containing wastewaters.

5. Conclusions
The dyes are natural and synthetic compounds that make the world more beautiful through
coloured products but are also considered as pollutants of some water resources.
The textile sector will continue to be vitally important in the area of water conservation due
to its high consumption of water resources, and its individual or combined effluents’
treatments for no environmental pollution generation (i.e. polluting colourants).
The satisfaction of both discharge criteria for sewerage systems, watercourses and textile
reuse standards within economically viable limits implies critical analyses of industrial
effluents (total wastewater and raw reusable stream characterisation) and removal of all
pollutants from final effluents. The special category of organic pollutants - textile organic
dyes - must respect the strict limits in final effluents discharged or not in natural water
resources. This fact imposes the colour and/or dye removal from final effluents (especially
industrial effluents).
Dye removal from textile effluents in controlled conditions and strict reproductibility is an
environmental issue achievable by application of adequate mechano-physico-chemical and
also biological treatment procedures.
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 83

6. References
Adams, C.D. & Gorg, S. (2002). Effect of pH and gas-phase ozone concentration on the
decolourization of common textile dyes. J.Environ.Eng., Vol.128, No.3, pp. 293-298
Alves, B.M.A. & Pinho, D.N.M. (2000). Ultrafiltration for colour removal of tannery dyeing
wastewaters., Desalination, Vol.1303, pp. 147-154
Anjaneyulu, Y.; Sreedhara Chary, N. & Suman Raj, D.S. (2005). Decolourization of industrial
effluents – available methods and emerging technologies – a review. Reviews in
Environmental Science and Bio/Technology, Vol.4, pp. 245-273, DOI 10.1007/s11157-
005-1246-z
Baban, A.; Yediler, A. & Ciliz, N.K. (2010). Integrated water management and CP
implementation for wool and textile blend processes. Clean, Vol.38, No.1, pp. 84-90
Balchioglu, I.A.; Aslan, I. & Sacan, M.T. (2001). Homogeneous and heterogeneous advanced
oxidation of two commercial reactive dyes. Environ.Technol., Vol.22, pp.813-822
Banat, M.E.; Nigam, P.; Singh, D. & Marchant, R. (1996) Microbial decolourization of textile
dye containing effluents, a review. Biores. Technol., Vol.58, pp. 217-227
Bhattacharyya, K.G. & Sarma, A. (2003). Adsorption characteristics of the dye, Brilliant
Green, on Neem leaf powder. Dyes Pigments, Vol.57, pp. 211-222
Birch, R.R.; Biver, C. Campagna, R.; Gledhill, W.E.; Pagga, U.; Steber, J.; Reust, H. &
Bontinck, W.J. (1989). Screening of chemicals for anaerobic biodegradability.
Chemosphere, Vol.19, No.(10-11), pp. 1527-1550
Bisschops, I.A.E. & Spanjers, H. (2003). Literature review on textile wastewater
characterisation. Environmental Technology, Vol.24, pp. 1399-1411
Bertea, A. & Bertea, A.P. (2008). Decolorisation and recycling of textile wastewater (in
Romanian), Performantica Ed, ISBN 978-973-730-465-0, Iasi, Romania
Börnick, H. & Schmidt, T.C. (2006). Amines, In: Organic pollutants in the water cycle. Properties,
occurrence, analysis and environmental relevance of polar compounds, T. Reemtsma & M.
Jekel, (Eds.), pp. 181-208, Wiley-VCH Verlag GmbH & Co. KGaA, ISBN 978-3-527-
31297-9, Weinheim, Germany
Borrely, S.I.; Cruz, A.C.; Del Mastro, N.L.; Sampa, M.H.O. & Somssari, E.S. (1998). Radiation
process of sewage and sludge – A review. Prog.Nucl.Energy, Vol.33, No.1/2, pp. 3-21
Carliell, C.M.; Barclay, S.J.; Naidoo, N.; Buckley, C.A.; Mulholland, D.A. & and Senior, E.
(1994). Anaerobic decolorisation of reactive dyes in conventional sewage treatment
processes. Water SA, Vol.20, pp. 341-344
Ciardelli, G. & & Ranieri, N. (2001). The treatment and reuse of wastewater in the textile
industry by means of ozonation and electroflocculation. Water Res., Vol.35, pp. 567-
572
Cooper P. (1995). Color in dyehouse effluent, Society of Dyers and Colourists, ISBN 0 901956
694, West Yorkshire BDI 2JB, England
Correia, V.M.; Stephenson, T. & Judd, S.J. (1994). Characterization of textile wastewaters – a
review. Environmental Technology, Vol.15, pp.917-929
Daneshvar, N.; Sorkhabi, H.A. & Tizpar, A. (2003). Decolorization of orange II by
electrocoagulation method. Sep.Purifi.Technol., Vol.31, pp.153-162
Dascalu T.; Acosta-Ortiz, E.S.; Morales, O.M. & Compean, I. (2000). Removal of the indigo
colour by laser beam-denim interanction. Opt. Lasers Eng., Vol.34, pp.179-189
Dos Santos, A.B.; Cervantes, F.J. & Van Lier, J.B. (2004). Azo dye reduction by thermophilic
anaerobic granular sludge, and the impact of the redox mediator AQDS on the
reductive biochemical transformation. Applied Microbiology and Biotechnology,
Vol.64, pp.62-69
Organic Pollutants Ten Years After
84 the Stockholm Convention – Environmental and Analytical Update

EPA. (1997). Profile of the textile industry. Environmental Protection Agency, Washington, USA
EWA. (2005). Efficient use of water in the textile finishing industry, Official Publication of the
European Water Association (EWA), Brussels, Belgium
Fontenot, E.J.; Lee, Y.H.; Matthews, R.D.; Zhu, G. & Pavlostathis, S.G. (2003). Reductive
decolorization of a textile reactive dyebath under methanogenic conditions. Applied
Biochemistry and Biotechnology, Vol.109, pp. 207-225
Ghayeni, S.B.; Beatson, P.J.; Schneider, R.P. & Fane, A.G. (1998). Water reclamation from
municipal wastewater using combined microfiltration-reverse osmosis (ME-RO):
Preliminary performance data and microbiological aspects of system operation.
Desalination, Vol.116, pp. 65-80
Gupta, G.S.; Singh, A.K.; Tayagi, B.S.; Prasad, G. & Singh, V.N. (1992). Treatment of carpet
and metallic effluent by China clay. J.Chem.Technol.Biotechnol., Vol.55, pp. 227-283
Hao, O.J.; Kim, H. & Chang, P.C. (2000). Decolorization of wastewater. Critical Reviews in
Environmental Science and Technology, Vol.30, pp. 449-505
Kim, T.H.; Park, C.; Lee, J.; Shin, E.B. & Kim, S. (2002). Pilot scale treatment of textile
wastewater by combined processes (fluidized biofilm process– chemical
coagulation – electrochemical oxidation). J.Hazar.Mat. B, Vol.112, pp.95-103
Krapfenbauer, K.F.; Robinson, M.R. & Getoff, N. (1999). Development of and testing of TiO2-
Catalysts for EDTA-radiolysis using γ-rays (1st part). J.Adv.Oxid.Technol., Vol.4,
No.2, pp.213-217
Lin, S.H. & Chen, L.M. (1997). Treatment of textile waste waters by chemical methods for
reuse. Water Res., Vol.31, No.4, pp.868-876
Lopez, A.; Ricco, G.; Ciannarella, R.; Rozzi, A., Di Pinto, A.C. & Possino, R. (1999). Textile
wastewater reuse: ozonation of membrane concentrated secondary effluent. Water
Sci. Technol., Vol.40, pp. 99-105
Machenbach, I. (1998). Membrane technology for dyehouse effluent treatment. Membrane
Technol., Vol.58, pp.7-10
Macoveanu, M.; Bilba, D.; Bilba, N.; Gavrilescu, M. & Soreanu, G. (2002). Ionic exchange
processes in environmental protection (in Romanian), MatrixRom Ed., Bucuresti,
Romania
Matsui, Y.; Murase, R.; Sanogawa, T.; Aoki, N.; Mima, S.; Inoue, T. & Matsushita, T. (2005).
Rapid adsorption pretreatment with submicrometre powdered activated carbon
particles before microfiltration. Water Science and Technology, Vol.51, pp.249-256
Mignani, M.G.; Nosenzo, G. & Gualdi, A. (1999). Innovative ultrafiltration for wastewater
reuse. Desalination, Vol.124, pp.287-292
Naveed, S.; Bhatti, I. & Ali, K. (2006). Membrane technology and its suitability for treatment
of textile waste water in Pakistan. Journal of Research (Science), Bahauddin Zakariya
University, Multan, Pakistan, Vol. 17, No. 3, pp.155-164
Neamtu, M.; Zaharia, C.; Catrinescu, C.; Yediler, A.; Kettrup, A. & Macoveanu, M. (2004).
Fe-exchanged Y zeolite as catalyst for wet peroxide oxidation of reactive azo dye
Procion Marine H-EXL. Applied Catalysis B: Environmental, Vol.78, No.2, pp.287-294
Nigam, P.; Armou, G.; Banat, I.M.; Singh, D. & Marchant, R. (2000). Physical removal of
textile dyes and solid-state fermentation of dye-adsorbed agricultural residues.
Biores. Technol., Vol.72, pp.219-226
Omura, T. (1994). Design of chlorine – fast reactive dyes – part 4; degradation of amino
containing azo dyes by sodium hydrochlorite. Dyes Pigments, Vol.26, pp.33-38
Ong, S.A.; Toorisaka, E.; Hirata, M. & Hano, T. (2005). Treatment of azo dye Orange II in
aerobic and anaerobic-SBR systems. Proc.Biochem., Vol.40, pp.2907-2914
Textile Organic Dyes – Characteristics, Polluting Effects
and Separation/Elimination Procedures from Industrial Effluents – A Critical Overview 85

Orhon, D.; Babuna, F.G. & Insel, G. (2001). Characterization and modelling of denim-
processing wastewaters for activated sludge. Journal of Chemical Technology and
Biotechnology, Vol.76, pp.919-931
Ozcan, A.S.; Erdem, B. & Ozcan, A. (2004). Adsorption of Acid Blue 193 from aqueous
solutions onto Na-bentonite and DTMA-bentonite. J.Coll.Interf.Sci., Vol.280, No.1,
pp.44-54
Oztekin, Y; Yazicigil, Z.; Ata, N. & Karadayl, N. (2010). The comparison of two different
electro-membrane processes‘ performance for industrial application. Clean-Soil, Air,
Water, Vol.38, No.5-6, pp.478-484
Ramesh Babu, B.; Parande, A.K.; Raghu, S. & Prem Kumar, T. (2007). Textile technology.
Cotton Textile Processing: Waste Generation and Effluent Treatment. The Journal of
Cotton Science, Vol.11, pp.141-153
Robinson, T.; McMullan, G.; Marchant, R. & Nigam, P. (2001). Remediation of dyes in textile
effluent: a critical review on current treatment technologies with a proposed
alternative. Bioresource Technology, vol.77, pp.247-255
Slokar, Y.M. & Le Marechal, A.M. (1997). Methods of decoloration of textile wastewaters.
Dyes Pigments, Vol.37, pp.335-356
Sarasa, J.; Roche, M.P.; Ormad, M.P.; Gimeno, E.; Puig, A. & Ovelleiro, J.L. (1998). Treatment
of wastewater resulting from dye manufacturing with ozone and chemical
coagulation. Water Res., Vol.32, No.9, pp.2721-2727
Soloman, P.A.; Basha, C.A.; Ramamurthi, V.; Koteeswaran, K. & Balasubramanian, N.
(2009). Electrochemical degradation of Remazol Black B dye effluent. Clean, Vol.37,
No.11, pp.889-900
Stolz, A. (2001). Basic and applied aspects in the microbial degradation of azo dyes.
Appl.Microbiol.Biotechnol., Vol.56, pp.69-80
Surpăţeanu, M. & Zaharia, C. (2004a). Advanced oxidation processes. Decolorization of
some organic dyes with hydrogen peroxide. Environmental Engineering and
Management Journal, Vol.3, No.4, pp.629-640
Surpateanu, M. & Zaharia C. (2004b). Advanced oxidation processes for decolorization of
aqueous solution containing Acid Red G azo dye. Central European Journal of
Chemistry, Vol.2, No.4, pp.573-588
Suteu, D.; Bîlbă, D. & Zaharia, C. (2002). Kinetic study of Blue M-EB dye sorption on ion
exchange resins. Hungarian Journal of Industrial Chemistry, Vol.30, pp.7-11
Suteu, D.; Zaharia, C.; Bilba, D.; Muresan, A.; Muresan, R. & Popescu, A. (2009a).
Decolorization wastewaters from the textile industry – physical methods, chemical
methods. Industria Textila, Vol.60, No.5, pp.254-263
Suteu, D.; Zaharia, C.; Muresan, A.; Muresan, R. & Popescu, A. (2009b). Using of industrial
waste materials for textile wastewater treatment. Environmental Engineering and
Management Journal, Vol.8, No.5, pp.1097-1102
Suteu, D. & Zaharia, C. (2008). Removal of textile reactive dye Brilliant Red HE-3B onto
materials based on lime and coal ash, ITC&DC, Book of Proceedings of 4th
International Textile, Clothing & Design Conference – Magic World of Textiles, pp.1118-
1123, ISBN 978-953-7105-26-6, Dubrovnik, Croatia, October 5-8, 2008
Suteu, D.; Zaharia, C. & Malutan, T. (2011a). Biosorbents Based On Lignin Used In
Biosorption Processes From Wastewater Treatment (chapter 7). In: Lignin: Properties
and Applications in Biotechnology and Bioenergy, Ryan J. Paterson (Ed.), Nova Science
Publishers, 27 pp., ISBN 978-1-61122-907-3, New York, U.S.A.
Organic Pollutants Ten Years After
86 the Stockholm Convention – Environmental and Analytical Update

Suteu, D.; Zaharia, C. & Malutan, T. (2011b). Removal of Orange 16 reactive dye from
aqueous solution by wasted sunflower seed shells. Journal of the Serbian Chemical
Society, Vol.178, No.3, pp.907-924
Tan, N.C.G.; Borger, A.; Slenders, P., Svitelskaya, A.; Lettinnga, G. & Field, J.A. (2000).
Degradation of azo dye mordent yellow 10 in a subsequential anaerobic and
bioaugmented aerobic bioreactor. Water Sci.Technol., Vol.42, No.(5-6), pp.337-344
Tang, C. & Chen, V. (2002). Nanofiltration of textile wastewater for water reuse. Desalination,
Vol.143, pp.11-20
Vandevivere, p.C.; Bianchi, R. & Verstraete, W. (1998). Treatment and reuse of wastewater
from the textile wet-processing industry review of emerging technologies.
J.Chem.Technol.Biotechnol., Vol.72, No.4, pp.289-302
Venkat Mohan, s.; Srimurli, M.; Sailaja, P. & Karthikeyan, J. (1999). A study of acid dye
colour removal using adsorption and coagulation. Environ.Eng.Poly., vol.1, pp.149-
154
Vlyssides, A.G.; Papaioannou, D.; Loizidoy, M.; Karlis, P.K. & Zorpas, A.A. (2000). Testing
an electrochemical method fortreatment of textile dye wastewater. Waste Manag.,
Vol.20, pp.569-574
Xu, Y. & Lebrun, R.E. (1991). Treatment of textile dye plant effluent by nanofiltration
membrane. Separ.Sci.Technol., Vol.34, pp.2501-2519
Welham, A. (2000). The theory of dyeing (and the secret of life). Journal of the Society of Dyers
and Colourists, Vol.116, pp.140-143
Wiesmann, U.; Choi, I.S. & Dombrowski, E.M. (2007). Fundamentals of Biological Wastewater
Treatment. Wiley-VCH Verlag GmbH&Co. KgaA, Weinheim, Germany
Yang, Y.; Wyatt II, D.T & Bahorsky, M. (1998). Decolorisation of dyes using UV/H2O2
photochemical oxidation. Text.Chem.Color., Vol.30, pp.27-35
Zaharia, C. (2006). Chemical wastewater treatment (in Romanian), Performantica Ed., ISBN
978-973-730-222-9, Iasi, Romania
Zaharia, C. (2008). Legislation for environmental protection (in Romanian), Politehnium Ed.,
ISBN 978-973-621-219-2, Iasi, Romania
Zaharia, C.; Diaconescu, R. & Surpăţeanu, M. (2006). Optimization study of a wastewater
chemical treatment with PONILIT GT-2 anionic polyelectrolyte. Environmental
Engineering and Management Journal, Vol.5, No.5, pp.1141-1152
Zaharia, C.; Diaconescu, R. & Surpăţeanu, M. (2007). Study of flocculation with Ponilit GT-2
anionic polyelectrolyte applied into a chemical wastewater treatment. Central
European Journal of Chemistry, Vol.5, No.1, pp.239-256
Zaharia, C.; Surpăţeanu, M.; Creţescu, I.; Macoveanu, M. & Braunstein, H. (2005).
Electrocoagulation/electroflotation – methods applied for wastewater treatment.
Environmental Engineering and Management Journal, Vol.4, No.4, pp.463-472
Zaharia, C.; Suteu, C. & Muresan, A. (2011). Options and solutions of textile effluent
decolourization using some specific physico-chemical treatment steps. Proceedings
of 6th International Conference on Environmental Engineering and Management
ICEEM’06, pp. 121-122, Balaton Lake, Hungary, September 1-4, 2011
Zaharia, C.; Suteu, D.; Muresan, A.; Muresan, R. & Popescu, A. (2009). Textile wastewater
treatment by homogenous oxidation with hydrogen peroxide. Environmental
Engineering and Management Journal, Vol.8, No.6, pp.1359-1369

View publication stats

You might also like